Open access peer-reviewed chapter

Mechanisms of Resistance to Quinolones

Written By

Sandra Georgina Solano-Gálvez, María Fernanda Valencia-Segrove, María José Ostos Prado, Ana Berenice López Boucieguez, Diego Abelardo Álvarez-Hernández and Rosalino Vázquez-López

Submitted: 27 September 2019 Reviewed: 16 April 2020 Published: 03 July 2020

DOI: 10.5772/intechopen.92577

From the Edited Volume

Antimicrobial Resistance - A One Health Perspective

Edited by Mihai Mareș, Swee Hua Erin Lim, Kok-Song Lai and Romeo-Teodor Cristina

Chapter metrics overview

1,807 Chapter Downloads

View Full Metrics

Abstract

Antimicrobial resistance is a worldwide problem. Various pathogenic bacteria can be resistant to one or several antibiotics, resulting in a serious public health problem. Isolation of pathogenic bacteria resistant to multiple last-generation antibiotics from hospital samples have been reported. In that sense, the isolation of pathogenic strains resistant to members of the quinolone family, from clinical samples, is an increasing phenomenon. Quinolones are a group of synthetic broad-spectrum antimicrobials, whose mechanism of action is the inhibition of DNA gyrase and topoisomerase IV, with the consequent DNA breakdown and cell death due to genotoxic damage. Three mechanisms have been determined by which bacteria can be resistant to quinolones: (1) Chromosomal mutations in coding genes (mutations that alter the objectives of the drug). (2) Mutations associated with the reduction of the intracytoplasmic concentration of quinolones. (3) Plasmid-mediated quinolone resistance genes (plasmids that protect cells from the lethal effects of quinolones). In this chapter, we analyze each of them and provide the most current connections and investigations of these processes.

Keywords

  • antibiotic resistance
  • quinolones
  • fluoroquinolones
  • DNA topoisomerase IV
  • genotoxic damage

1. Background

Antimicrobial resistance has become a serious public health problem in recent years. This problem has been increasing and is currently a truly global crisis that offers one of the worst forecasts of catastrophic scenarios in public health worldwide.

A sign of the seriousness of the problem is the fact that World Health Organization (WHO)'s new Global Antimicrobial Surveillance System (GLASS) reported the widespread occurrence of antibiotic resistance among 500,000 people with suspected bacterial infections across 22 countries [1].

Likewise, Centers for Disease Control (CDC)‘s Antibiotic Resistance Threats in the United States (US), in 2019 (2019 AR Threats Report), reported that more than 2.8 million antibiotic-resistant infections occur in the US each year, and more than 35,000 people die as a result. Besides, 223,900 cases of Clostridium difficile occurred in 2017 and at least 12,800 people died [2].

Many bacteria produce important infections in human health, either due to community-acquired infections, nosocomial infections, or at intensive care units. Among these, many have an important phenotypic profile of antibiotic resistance. For example, Staphylococcus aureus, Enterococcus spp., Enterobacteriaceae (other than Salmonella and Shigella), Pseudomonas aeruginosa, and Acinetobacter spp. [3, 4].

To classify these microorganisms according to the degree of resistance and acquired resistance profiles, a group of experts in the field of antimicrobial resistance in joint work with the European Center for the Prevention of Diseases and Control (ECDC) and the CDC established the definitions and characteristics among resistant bacteria: multidrug-resistant (MDR), extensively drug-resistant (XDR), and pan drug-resistant (PDR) bacteria [3, 4].

To establish objective parameters of the phenotypic resistance profile in each of these bacteria, epidemiologically significant antimicrobial categories were established. These categories were established based on the documents and cut-off points of the Clinical Laboratory Standards Institute (CLSI), the European Antimicrobial Sensitivity Testing Committee (EUCAST), and the US Food and Drug Administration (FDA) [3, 4].

Based on the new limits and definitions: MDR bacteria possess acquired resistance to at least one antibiotic of three or more categories; XDR bacteria possess resistance to at least one antibiotic of almost all categories, except one or two of them; and PDR bacteria are resistant to all agents of all categories of antimicrobials [3, 4].

Antimicrobial resistance has been observed in all families of antibiotics, including the latest generation and intrahospital antibiotics such as quinolones.

The wide use of quinolones in clinical practice includes the administration of the antibiotic in prophylaxis, in neutropenic patients with cancers, in cirrhotic patients at risk for spontaneous bacterial peritonitis, and in urologic surgery, among others. In many of these cases, strains with varying degrees of resistance to quinolones have been isolated [5, 6].

Advertisement

2. History

In 1962, quinolones were discovered as an important treatment for various pathological manifestations. The first one was nalidixic acid, which was synthetically produced by George Lesher at the Sterling-Winthrop Research Institute. It was synthesized from the isolation of chloro-1-ethyl-1,4-dihydro-4-oxo-3-quinoline carboxylic acid years before, as a product derived from the synthesis of chloroquine [7]. Its origin dates back to the use of chloroquine as an antimalarial agent. It was until years after its development that nalidixic acid was approved for the treatment of urinary tract infections by Gram-negative bacteria. This compound does not have an important effect on Gram-positive bacteria, in addition to having a certain cytotoxic effect on the gastrointestinal tract and the central nervous system. Its effect on Gram-negative bacteria is characteristic of the first generation of quinolones [8].

Advertisement

3. Epidemiology

The indiscriminate prescription of quinolones worldwide has led to a rapid increase in bacterial resistance. Acinetobacter spp., Campylobacter spp., Capnocytophaga spp., Clostridium spp., Escherichia coli, Klebsiella pneumoniae, Mycobacterium tuberculosis, Neisseria gonorrhea, Proteus mirabilis, P. aeruginosa, Salmonella spp., S. aureus, and Streptococcus pneumoniae, among others, have been reported as resistant [7, 9, 10].

The ECDC collects and reports through the European Antimicrobial Resistance Surveillance Network (EARS-Net) information of seven bacterial pathogens that commonly cause infections in humans: Acinetobacter spp., Enterococcus faecalis, Enterococcus faecium, Escherichia coli, Klebsiella pneumoniae, Staphylococcus aureus, Streptococcus pneumoniae, and Pseudomonas aeruginosa. Information comparing their profile of resistance to quinolones in Europe between 2015 and 2018 can be found in Table 1 [11, 12].

Bacteria Mean resistance percentage Country with the lowest resistance percentage Country with the highest resistance percentage
2015 2018 2015 2018 2015 2018
Acinetobacter spp. 43.7% 36.2% Belgium (0%) Norway (0%) Greece (94.9%) Croatia (96.1%)
Escherichia coli 22.8% 25.3% Iceland (6.8%) Finland (11.4%) Cyprus (45.5%) Cyprus (42.4%)
Klebsiella pneumoniae 29.7% 31.6% Iceland (2.9%) Iceland (0%) Slovakia (70%) Poland (68.2%)
Pseudomonas aeruginosa 19.3% 19.7% Estonia (0%) Malta (0%) Romania (59%) Slovakia (52.4%)

Table 1.

Profile of resistance to quinolones of European countries (2015 vs. 2018).

The EARS-Net report does not contain information about quinolone resistance to other bacteria. Adapted from: EARS-Net 2015 and Ecdc. SURVEILLANCE REPORT. 2018 [11, 12].


Advertisement

4. The structure of quinolones

The structure of quinolones derives from two types of rings, a naphthyridine core with a nitrogen molecule in positions 1 and 8. Through this structure, the compound is limited to being used as a therapy against Gram-negative bacteria. However, it has been shown that by inserting a cyclopropyl group in the first position of the nitrogen ring, an effect is achieved not only on Gram-negative bacteria but also on Gram-positive ones (Figure 1) [13, 14].

Figure 1.

Molecular structure of representative members of each quinolone generation. Based on PubChem public archive https://www.ncbi.nlm.nih.gov/pcsubstance [14, 17, 19, 20].

The second generation was developed in 1980, from the addition of a fluorine atom at position six, resulting in fluoroquinolones. These have higher activity in Gram-negative bacteria, as well as Gram-positive bacteria. Some fluoroquinolones can inhibit all Gram-negative organisms. Quinolones with piperazine on carbon 7 are effective in Gram-negative bacteria and the signaling of topoisomerase 4 (Figure 1) [13, 15, 16, 17].

Later, the third generation arises by adding certain molecules in the rings, such as the cyclopropyl ring in the first position of nitrogen, improving the activity in Gram-positive bacteria. Some of these modifications achieved sensitivity in organism resistant to different antibiotics, including Streptococcus pneumoniae. Other benefits of this generation are a longer life in serum and activity against anaerobic organisms (Figure 1) [7, 18, 19].

The fourth generation was later developed by incorporating nitrogen in the eighth position, resulting in a broad-spectrum antibiotic. Its action in some Gram-positive organisms is more effective compared to the other generations; however, its activity in anaerobic organisms is limited. It has a superior bacterial selectivity to avoid a high level of resistance and its toxic effects are less unfavorable than in the other generations [7, 8]. Thanks to the modifications made to the quinolones, an improvement in its pharmacokinetics and pharmacodynamics has been obtained, thus optimizing absorption, metabolism, and elimination, achieving lower toxicity and superiority in the mechanisms of action. It has also been possible to modify the half-life of the drug making only one dose per day necessary (Figure 1) [20].

Currently, nine fluoroquinolones have been approved in the US while others continue to be used in clinical trials. Information regarding the generations, compounds, and spectrum of activity can be found in Table 2.

Generation Compounds Activity spectrum
1 Nalidixic acid Gram-negative bacteria (not Pseudomonas spp.)
2 2a Ciprofloxacin, enoxacin, norfloxacin Gram-negative bacteria and atypical pathogens (Mycoplasma pneumoniae and Chlamydia pneumoniae)
2b Levofloxacin, lomefloxacin, ofloxacin Gram-negative bacteria, Gram-positive bacteria (not Streptococcus pneumoniae), and atypical pathogens
3 Clinafloxacin, gatifloxacin, grepafloxacin, sparfloxacin. Gram-negative bacteria, Gram-positive bacteria (Streptococcus pneumoniae) and improved activity against atypical pathogens
4 Gatifloxacin, gemifloxacin, moxifloxacin, trovafloxacinin Gram-negative bacteria, Gram-positive bacteria (Streptococcus pneumoniae) and improved activity against atypical and anaerobic pathogens

Table 2.

Classification of quinolones.

Adapted from: Pham TDM, Ziora ZM, Blaskovich MAT [7].

It has been reported that several agents such as Escherichia coli, Klebsiella pneumoniae, Neisseria gonorrhoeae, or Staphylococcus aureus have presented significant resistance to quinolones [21]. Plasmid-mediated quinolone resistance (PMQR) was a completely unexpected event since it was thought that the only mutation would occur in genes encoding topoisomerase II identification. Currently, the resistance mechanism is multifactorial. However, the most common quinolone resistance mechanism is topoisomerase mutations [15, 22].

The excessive use of this type of drug has caused the incidence rates of hypersensitivity to increase more and more, taking the second place of antibiotics with a greater number of hypersensitivity reactions in in-hospital patients. The main agents that cause hypersensitivity are ciprofloxacin, levofloxacin, and moxifloxacin. This has positioned quinolones as the non-beta-lactam antibiotics with the highest incidence of hypersensitivity reactions [23, 24].

Advertisement

5. Mechanism of action

The mechanism of action of quinolones is based on the inhibition of bacterial topoisomerases II and IV. Topoisomerases are enzymes responsible for maintaining the tertiary structure of DNA during various cellular processes, such as synthesis, replication, condensation, and decondensation of DNA, among others [25, 26, 27, 28, 29].

Topoisomerase II, also known as DNA gyrase, is considered a negative supercoiling enzyme, which means that it cuts the two strands of DNA and propitiates that the DNA is twisted to the left producing a twist in a way contrary to the direction of the double helix. This enzyme participates in the DNA winding and relaxation during various processes, mainly in the synthesis and replication of DNA [30, 31].

The DNA gyrase consists of a heterotetramer, which is formed by two GyrA subunits and two GyrB subunits. The GyrA subunits participate in the union with the DNA and are responsible for making the double helix cuts. The GyrB subunits possess ATPase activity [30].

Topoisomerase IV is responsible for preventing the chromatids from being chained, meaning it participates in the separation of daughter chromatids after DNA replication [32].

Like DNA gyrase, topoisomerase IV is made up of a tetramer. It has two ParC subunits and two ParE subunits. These subunits possess homologous activity of GyrA and GyrB, respectively [32].

When quinolones interact and inhibit topoisomerase II and IV, it induces DNA breakdown and cell death due to genotoxic damage [27, 28, 29].

Advertisement

6. Resistance mechanisms

To counteract the effect of quinolones, bacteria have developed various resistance mechanisms to these antibiotics. Bacterial resistance to quinolones is mainly based on three points (Table 3, Figure 2):

  1. Chromosomal mutations in coding genes (mutations that alter the objectives of the drug).

  2. Mutations associated with the reduction of the intracytoplasmic concentration of quinolones.

  3. PMQR genes (plasmids that protect cells from the lethal effects of quinolones) [33].

Mechanism Description
Chromosomal mutations in coding genes Occurs due to errors in the replication of the genes encoding the GyrA subunits of DNA gyrase and ParC of topoisomerase IV
Mutations associated with the reduction of the intracytoplasmic concentration of quinolones Occurs due to mutations that lead to a decrease in the intracytoplasmic concentration of the antibiotic. It may happen through:
Overexpression of efflux pumps from the resistance-nodulation-cell division Both Reduction of the membrane permeability by downregulation of extra-membrane proteins
Plasmid-mediated quinolone resistance genes Occurs due to the activation of plasmid-mediated quinolone resistance genes. Among them are:
Qnr’s encode proteins that protect DNA gyrase and topoisomerase IV AAC(6′)-lb-cr acetylates quinolones with an appropriate amino nitrogen target QepA and OqxAB, which increase the outflow of quinolones through efflux pumps

Table 3.

Mechanisms of resistance to quinolones.

Adapted from: Álvarez-Hernández DA, Garza-Mayén GS, Vázquez-López R. Quinolones. Nowadays perspectives and mechanisms of resistance [34].

Figure 2.

Schematic representation of the mechanisms of bacterial resistance to quinolones. Based on Susana Correia et al. [22].

6.1 Chromosomal mutations in coding genes (mutations that alter the objectives of the drug)

The quinolone resistance associated with chromosomal mutations occurs due to errors in the replication of the genes encoding the GyrA subunits of DNA gyrase and ParC of topoisomerase IV [33, 35].

In the amino acid sequences of the GyrA and ParC subunits, there are specific regions that interact with the DNA. In these regions, there are conserved domains called quinolone resistance determining region (QRDR) [31, 35, 36, 37, 38, 39].

It is precisely in the sequences that code for each of the QRDR domains of the GyrA and ParC subunit genes, where such mutations occur [31, 35, 36, 37, 38, 39].

It has been reported that quinolone resistance may also occur due to mutations in the genes encoding the GyrB and ParE subunits; however, they do not occur so frequently and their clinical value appears to be very limited [35, 40, 41].

There is evidence that in Gram-negative bacteria, DNA gyrase turns out to be more susceptible to inhibition than topoisomerase IV. On the other hand, in Gram-positive bacteria, the opposite phenomenon occurs; that is, that topoisomerase IV is more susceptible to inhibition than gyrase. However, certain bacteria show the opposite effect, being the exception to the rule [31, 42, 43].

Therefore, we can affirm that the phenomenon of resistance in the majority of Gram-negative bacteria occurs mainly in GyrA, while in most Gram-positive bacteria the inhibition of ParC is the most important [31, 42, 43].

Summarizing, mutations that occur in the sequences encoding the QRDR domains in both GyrA-ParC and GyrB-ParE favor a decrease in the binding affinity of quinolones with the DNA–DNA gyrase and DNA-topoisomerase IV complex [33, 35].

6.2 Mutations associated with the reduction of the intracytoplasmic concentration of quinolones

Another important quinolone resistance mechanism consists in the ability of the bacteria to decrease the intracytoplasmic concentration of the antibiotic; this decrease in concentration is determined by certain mutations.

This phenomenon is achieved through three mechanisms:

  1. Efflux pumps that promote the active transport of quinolones to the outside of the bacterial cell.

  2. Decreased membrane permeability toward the antibiotic.

  3. A combination of both mechanisms.

It has been described that only efflux pumps participate in Gram-positive bacteria as mechanisms to reduce the intracytoplasmic concentration of quinolones since there is no evidence that the decrease in cytoplasmic membrane permeability participates in this type of bacteria [44].

On the other hand, Gram-negative bacteria do have both mechanisms and participate in a complementary way with one another, the decrease in permeability in the cytoplasmic membrane being the most important for these bacteria [45].

These two mechanisms involved in the decrease of the intracytoplasmic concentration of quinolones are not induced by the drugs themselves. There is evidence that these two mechanisms occur because of mutations in genes that encode regulatory proteins that control transcription of the outflow pump or genes that code for porin synthesis [35, 46].

6.2.1 Mutations associated with the reduction of the intracytoplasmic concentration of quinolones in Gram-positive bacteria

This resistance mechanism in Gram-positive bacteria is associated with the presence of chromosomally encoded efflux pumps that decrease the intracytoplasmic concentration of the antibiotic, giving the bacteria the characteristic of being MDR.

Efflux pumps are classified into two groups: primary active transporters and secondary active transporters [47].

The primary active transporter proteins are pumps that use ATP as a source of energy. This type of primary active transporter integrates the members of the ATP-binding cassette (ABC) superfamily [48, 49, 50].

On the other hand, the secondary active transporter proteins use the energy obtained by the difference of chemical gradients formed by either protons or ions, for example, sodium ions [48, 49].

Four types of secondary active transporter proteins have been identified: [47, 48, 49].

  1. The small multidrug-resistance (SMR) family

  2. The major facilitator superfamily (MFS)

  3. Multidrug and toxic compound extrusion (MATE) family

  4. The resistance-nodulation-cell division (RND) superfamily.

6.2.1.1 SMR (the small multidrug-resistance family)

Members of this family are proteins made up of an antiparallel dimer. Each monomer of this dimer has four transmembrane helices (TM1, TM2, TM3, and TM4). The TM 1 to M3 helices comprise the substrate binding pocket, while each TM4 helix is responsible for SMR TM4-TM4 dimerization [51, 52, 53].

The members of the SMR family are associated with resistance to various toxic compounds and some antibiotics; however, they do not appear to play a relevant role in resistance to quinolones.

6.2.1.2 MFS (the major facilitator superfamily)

Concerning efflux pumps related to the intracytoplasmic decrease in quinolone and consequently linked to resistance to this drug, they are efflux pumps that are part of the MFS. Three members of this family associated with quinolone resistance have been identified: NorA, NorB [50], and NorC [54]. Overexpression of each of three efflux pumps increases resistance to quinolones four to eight times [33].

6.2.1.2.1 NorA

The chromosomal gene that codes for NorA could be identified in 1986 from the isolation of Staphylococcus aureus obtained from a urine sample from a patient who had received treatment with norfloxacin at Teikyo University Hospital Japan [55]. It has been observed that NorA participates in the pumping of various quinolones, mainly ciprofloxacin and norfloxacin [56, 57].

Subsequent studies of genetic diversity described three alleles for the NorA gene [58]: NorAI (Yoshida), NorAII (Noguchi), and NorAIII (Kaatz). A correlation has been observed between the different types of NorA alleles and specific lineages of S. aureus. This fact suggests that there is a correlation between the NorA variants and the population structure (lineages) of this bacterium [58].

6.2.1.2.2 NorB

It has been described that the expression of the efflux pump NorB gives certain bacteria (e.g., Staphylococcus aureus) the adaptability in tissue infection conditions, even in the absence of antibiotics. This fact occurs because NorB gives Staphylococcus aureus the ability to eliminate antibacterial substances present in the abscess and produced as a defense mechanism by the host. In this way, NorB not only participates in the quinolone resistance mechanism but also contributes to the pathophysiology of certain infections [59].

6.2.1.2.3 NorC

The efflux pump Norc enhances the exit of quinolones such as ciprofloxacin, garenoxacin moxifloxacin, and sparfloxacin out of the bacterial cell. Its expression is regulated negatively by MgrA [54].

Many regulatory proteins participate in a complex regulatory process in the gene expression of NorA, NorB, and NorC. One of these regulatory proteins is MgrA, which shows the ability to bind to the NorA promoter region. The overexpression of MgrA causes the inhibition of the expression of NorA, NorB, and NorC, in the opposite, resistance to quinolones is associated with a low activity of MgrA and the consequent overproduction of NorA, NorB, and NorC that will promote a decrease in the intracytoplasmic concentration of the drug [54, 60, 61, 62].

There is evidence that MgrA activity could be determined by environmental conditions in which the bacterium is found. Acid conditions, oxidative, as well as the presence of iron, could alter the activity of MgrA and consequently the expression of NorA, NorB, and NorC and its effect on the pumping of quinolone and its concentration in the bacterial cytoplasm [35, 59, 63, 64, 65].

On the other hand, another transcriptional regulator, called NorG, which activates the expression of NorA and NorB but suppresses the expression of NorC, has been described. It is important to understand that the regulation of the gene expression of NorA, NorB, and NorC results from a complex molecular framework where both activators and inhibitors participate and the balance between them, as well as the environmental and nutritional conditions in which the bacteria develops, will give as a result the resistance or the lack of it to quinolones [35, 61, 62, 66].

6.2.1.3 Other members of the MFS (major facilitator superfamily)

6.2.1.3.1 MdeA

MdeA gen was identified in an open reading frame (ORF) expression library of the S. aureus genome. The efflux pump protein MdeA belongs to the MFS using the proton motive force to energize the transport of its substrates [67, 68].

MdeA confers resistance to the biocides benzalkonium chloride, dequalinium, tetraphenylphosphonium, and to the dye ethidium bromide [67]. MdeA also confers resistance to multiple antibiotics among which are fusidic acid, mupirocin, novobiocin, and virginiamycin, and to some extent toward ciprofloxacin and norfloxacin [67, 68].

6.2.1.3.2 SdrM

In 2006 Yamada et al. cloned a new gene called SA1972 isolated from Staphylococcus aureus. The product obtained was called SdrM and it was proven that it conferred resistance to the bacteria against, acriflavine, ethidium bromide, and norfloxacin. SdrM was classified as an efflux pump belonging to the MFS [69].

6.2.1.3.3 QacB (III)

The qacA and qacB genes that code for efflux pump proteins (QacA and QacB, respectively) are present in methicillin-resistant Staphylococcus aureus (MRSA). The efflux pump QacA has two isoforms, while the pump QacB has four known as QacBI, QacBII, QacBIII, and QacBIV. It has been observed that the QacBIII variant confers resistance to S. aureus to fluoroquinolones [70].

6.2.1.4 MATE (multidrug and toxic compound extrusion family)

6.2.1.4.1 MepA

The efflux pump MepA belongs to the multidrug and toxic compound extrusion (MATE) family. MepA gives the bacterium a phenotypic MDR profile associated with low-level resistance to some quaternary ammonium compounds. It also confers resistance to certain antibiotics, mainly toward glycylcyclines and to a lesser extent resistance to ciprofloxacin and norfloxacin [71, 72, 73].

In addition to the efflux pump described above, there are other transporters in Gram-positive bacteria that participate in the decrease in the intracytoplasmic concentration of quinolones in the bacterial cell, participating in resistance to this drug. Some of these transporters are LmrS, Bmr, Bmr3 and Blt, PmrA66, LmrP67, PatAB69, SatAB70, LmrA71, FepA, FepR, and TetR [35].

6.2.2 Mutations associated with the reduction of concentration in Gram-negative bacteria

6.2.2.1 RND (resistance-nodulation-cell division superfamily)

Gram-negative bacteria use efflux pumps belonging to the RND superfamily as the main mechanism of resistance to quinolones. The efflux pump RND pumps are a molecular complex consisting of three elements (Figure 3) [49, 74, 75, 76, 77]:

  1. In the inner membrane is RND pump protein.

  2. An adapter protein from the MFP (membrane fusion protein) family located in the periplasmic space.

  3. In the outer membrane is an outer membrane channel protein (OMP) belonging to the outer membrane factor (OMF) family.

Figure 3.

Schematic representation of molecular structure of RND (resistance-nodulation-cell division superfamily). Based on Eun-Hae Kim et al. [77].

The adapter protein MFP links the pump RND and the OMF protein [49, 74, 75, 76, 77].

In E. coli, the presence of five RND efflux transporters has been reported:

  1. AcrAB [78, 79]

  2. AcrAD [80, 81]

  3. AcrEF [82]

  4. MdtABC [83, 84]

  5. MdtEF [85, 86]

6.2.2.2 AcrAB-TolC [acriflavine (Acr) efflux system]

The AcrAB-TolC or acriflavine (Acr) efflux system consists of three elements [75, 87]:

  1. The outer-membrane channel TolC

  2. In the periplasmic space is the AcrA protein, which bridges these two integral membrane proteins

  3. In the inner membrane is the secondary transporter AcrB.

There is evidence that the ratio between the proteins that make up this complex is 3: 6: 3, comprising an AcrB trimer, an AcrA hexamer, and a TolC trimer [75, 87].

It has been shown that various dyes can be accommodated in the transmembrane domain of the Acr efflux system, as well as doxorubicin, minocycline, and quinolone molecules [88, 89].

6.2.2.2.1 AcrAD

AcrAD is an antibiotic efflux pump complex of the RND type. It provides resistance to aminoglycosides such as amikacin, gentamicin, and tobramycin. There is no known effect on quinolone resistance [80, 90].

6.2.2.2.2 AcrEF

AcrEF is an antibiotic efflux pump complex of the resistance-nodulation-cell division (RND) type. It provides resistance to cephalosporins, cephamycins, fluoroquinolones, and penams [91, 92].

6.2.2.2.3 MdtABC

MdtABC is an antibiotic efflux pump complex of the resistance-nodulation-cell division (RND) type. It provides resistance to aminocoumarins, which have a mechanism of action similar to quinolones [93, 94].

6.2.2.2.4 MdtEF

MdtEF is an antibiotic efflux pump complex of the RND type. It provides resistance to fluoroquinolones, macrolides, and penams [82].

6.2.2.3 Other members of the RND (resistance-nodulation-cell division superfamily)

6.2.2.3.1 MexAB-OprM efflux system

MexAB-OprM efflux system is an antibiotic efflux pump complex of the RND type. It provides resistance to multiple antibiotics, including aminocoumarins, carbapenems, cephalosporins, cephamycins, diaminopyrimidines, fluoroquinolones, macrolides, monobactams, penams, phenicols, peptides, sulfonamides, and tetracyclines [95, 96].

6.2.2.3.2 MexCD-OprJ with type A NfxB mutation

MexCD-OprJ with type A NfxB mutation is an antibiotic efflux pump complex of the RND type. It provides resistance to the aminocoumarins, cephalosporins, diaminopyrimidines, fluoroquinolones, macrolides, penams, phenicols, and tetracyclines [97].

6.2.2.3.3 MexCD-OprJ with type B NfxB mutation

MexCD-OprJ with type B NfxB mutation is an antibiotic efflux pump complex of the RND type. It provides resistance to the aminocoumarins, aminoglycosides, cephalosporins, diaminopyrimidines, fluoroquinolones, macrolides, penams, phenicols, and tetracyclines [97].

6.2.2.3.4 MexEF-OprN

MexEF-OprN is an antibiotic efflux pump complex RND. It provides resistance to diaminopyrimidines, fluoroquinolones, and phenicols [98].

6.2.2.3.5 MexXY-OprM

MexXY-OprM is an antibiotic efflux pump complex RND. It provides resistance to the acridine dye, aminoglycosides, carbapenems, cephalosporins, cephamycins, fluoroquinolones, macrolides, penams, phenicols, and tetracyclines [96, 99, 100].

6.2.2.3.6 CmeABC

CmeABC is an antibiotic efflux pump complex RND. It provides resistance to cephalosporins, fluoroquinolones, fusidic acid, and macrolides [101, 102].

6.2.2.3.7 AdeIJK

AdeIJK is an antibiotic efflux pump complex RND. It provides resistance to carbapenems, cephalosporins, diaminopyrimidines, fluoroquinolones, lincosamides, macrolides, penems, phenicols, rifamycins, and tetracyclines [103].

6.2.2.3.8 AdeABC

AdeABC is an antibiotic efflux pump complex RND. It provides resistance to glycylcyclines and tetracyclines [104, 105].

6.2.2.3.9 AdeL

AdeL is an antibiotic efflux pump complex RND. It provides resistance to fluoroquinolones and tetracyclines [106].

6.2.2.3.10 SmeDEF

SmeDEF is an antibiotic efflux pump complex RND. It provides resistance to fluoroquinolones, macrolides, phenicols, and tetracyclines [107].

Other molecular complexes associated with decreasing the intracytoplasmic concentration of antibiotics in Gram-negative bacteria include:

6.2.2.4 Members of the MFS (major facilitator superfamily) in Gram-negative bacteria

6.2.2.4.1 EmrAB-TolC

EmrAB-TolC is an antibiotic efflux pump belonging to MFS. It provides resistance to fluoroquinolones [108].

6.2.2.4.2 MdfA

MdfA is an antibiotic efflux pump belonging to MFS. It provides resistance to benzalkonium chloride, fluoroquinolones, rhodamine, and tetracyclines [109, 110].

6.2.2.5 Other Gram-negative mechanisms

Other molecular complexes associated with decreasing the intracytoplasmic concentration of antibiotics in Gram-negative bacteria include:

6.2.2.5.1 Porin OprF

The OprF porin channel is permeable to quinolones and other antibiotics, promoting its outflow and decreasing intracytoplasmic concentration and consequently is a mechanism of antibiotic resistance for the bacteria [111, 112].

6.3 Plasmid-mediated quinolone resistance genes (plasmids that protect cells from the lethal effects of quinolones)

In 1998 at the University of Alabama, from the isolation of Klebsiella pneumoniae from a urine sample, Martinez et al. managed to identify a plasmid they named pMG252. They demonstrated that this plasmid induced bacterial resistance to fluoroquinines and nalidixic acid. This resistance phenomenon could be induced in a variety of bacteria deficient in outer-membrane porins. They also described that this plasmid promoted the acceleration of resistance development and its propagation. The gene responsible for this resistance was called qnr, later it became qnrA [113, 114].

In 2002, Tran and Jacoby, working with the qnr plasmid, managed to identify an integron-like environment upstream from qacEΔ1 and sulI. The product obtained from this gene was a 218-aa protein called QnrA. This protein belonging to the pentapeptide repeat family shared sequence homology with the immunity protein McbG. Previous studies suggested that McbG protects DNA gyrase from the action of various genotoxic chemicals [115].

Based on the mechanism of action of quinolones (the inhibition of topoisomerases I and IV) and the similarity of QnrA to McbG, Tran and Jacoby determined the ability of QnrA to induce resistance against quinolones by topoisomerase protection [115].

In 2005, two independent teams managed to determine the same activity as QnrA for two other proteins identified as QnrB [116] and QnrS [117].

Subsequent studies of the qnrA plasmid found that this plasmid was able to promote greater resistance than expected and that is how, in 2006, Ari Robicsek et al. discovered another mechanism of action of resistance to quinolones mediated by the enzymatic action of aminoglycoside acetyltransferase, AAC(6′)-Ib-cr. They also reported that the quinolone resistance mechanism was determined by reduction of the activity of ciprofloxacin by N-acetylation at the amino nitrogen on its piperazinyl substituent [118].

In 2007, three groups of researchers separately demonstrated another resistance mechanism encoded by plasmids. These works, in correlation with Martinez’s works, involve quinolone efflux pumps mediated by plasmids QepA [119, 120] and OqxAB [121].

In summary, there are three mechanisms for PMQR:

  1. The plasmid genes qnrA, qnrB, qnrC, qnrD, qnrS, and qnrVC encode proteins from the pentapeptide repeat family that protects DNA gyrase and topoisomerase IV from quinolone inhibition. The qnr genes are generally associated with mobilizing or transposable elements in plasmids and are often incorporated into sul1-type integrons.

  2. The second mechanism mediated by plasmids involves acetylation of quinolones with an appropriate amino nitrogen target by a variant of the common aminoglycoside acetyltransferase AAC(6′)-Ib-cr.

  3. Improved outflow produced by plasmid genes for QepAB and OqxAB pumps.

Advertisement

7. Concluding remarks

Bacterial resistance to antibiotics is a serious problem worldwide and offers the bleakest outlook and prognosis. The number of reports of isolation of multiresistant strains is increasing, including antibiotics of the latest generation or exclusive intrahospital use. In this sense, isolates of strains resistant to practically all members of the quinolone family have been reported.

The implementation of appropriate practices in the use of antibiotics plays an important role in the fight against this serious global problem. The proper management of antibiotics must include limiting their use in the livestock, agricultural, and food industries; as well as the correct medical prescription, avoiding self-medication, and always seeking adherence to the full antibiotical treatment scheme.

The knowledge of the molecular mechanisms associated with resistance to quinolones and other antibiotics offers us great possibilities for molecular epidemiological monitoring of the emergence of new resistant strains, as well as their distribution. This knowledge offers the pharmaceutical industry the tools for the development of new drugs. It is important to consider that the development time of new drugs is exceeded by the speed of appearance of new resistant strains.

References

  1. 1. WHO. WHO | High Levels of Antibiotic Resistance Found Worldwide, New Data Shows [Internet] [cited 01 February 2020]. Available from: https://www.who.int/mediacentre/news/releases/2018/antibiotic-resistance-found/en/
  2. 2. Biggest Threats and Data | Antibiotic/Antimicrobial Resistance | CDC 2020 [Internet] [cited 01 February 2020]. Available from: https://www.cdc.gov/drugresistance/biggest-threats.html
  3. 3. Magiorakos A-P, Srinivasan A, Carey RB, Carmeli Y, Falagas ME, Giske CG, et al. Multidrug-resistant, extensively drug-resistant and pandrug-resistant bacteria: An international expert proposal for interim standard definitions for acquired resistance. Clinical Microbiology and Infection [Internet]. 2012;18(3):268-281 [cited 11 June 2019]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/21793988
  4. 4. Vázquez-López R, Rivero Rojas O, Ibarra Moreno A, Urrutia Favila JE, Peña Barreto A, Ortega Ortuño GL, et al. Antibiotic-resistant septicemia in pediatric oncology patients associated with post-therapeutic neutropenic fever. Antibiotics [Internet]. 2019;8(3):106. Available from: https://www.mdpi.com/2079-6382/8/3/106
  5. 5. Malekzadegan Y, Rastegar E, Moradi M, Heidari H, Sedigh E-SH. Prevalence of quinolone-resistant uropathogenic Escherichia coli in a tertiary care hospital in South Iran. Infection and Drug Resistance. 2019;12:1683-1689
  6. 6. Kim ES, Hooper DC. Clinical importance and epidemiology of quinolone resistance. Infection and Chemotherapy. 2014;46:226-238
  7. 7. TDM P, Ziora ZM, MAT B. Quinolone antibiotics. MedChemComm. 2019;10:1719-1739
  8. 8. Emmerson AM. The quinolones: Decades of development and use. The Journal of Antimicrobial Chemotherapy. 2003;51(90001):13-20
  9. 9. Piddock LJV. Fluoroquinolone resistance. British Medical Journal. 1998;317:1029-1030
  10. 10. Redgrave LS, Sutton SB, Webber MA, Piddock LJV. Fluoroquinolone resistance: Mechanisms, impact on bacteria, and role in evolutionary success. Trends in Microbiology. 2014;22:438-445
  11. 11. EARS-Net. Antimicrobial resistance surveillance in Europe; 2015
  12. 12. Ecdc. SURVEILLANCE REPORT. Surveillance of antimicrobial resistance in Europe 2018 [Internet]. 2018 [cited 07 February 2020]. Available from: www.ecdc.europa.eu
  13. 13. Appelbaum PC, Hunter PA. The fluoroquinolone antibacterials: Past, present and future perspectives. International Journal of Antimicrobial Agents. 2000;16:5-15
  14. 14. Nalidixic Acid—PubChem [Internet] [cited 12 February 2020]. Available from: https://pubchem.ncbi.nlm.nih.gov/substance/144075330
  15. 15. Aldred KJ, Kerns RJ, Osheroff N. Mechanism of quinolone action and resistance. Biochemistry. 2014;53:1565-1574
  16. 16. Kocsis B, Domokos J, Szabo D. Chemical structure and pharmacokinetics of novel quinolone agents represented by avarofloxacin, delafloxacin, finafloxacin, zabofloxacin and nemonoxacin. Annals of Clinical Microbiology and Antimicrobials. 2016b;15(1). DOI: 10.1186/s12941-016-0150-4
  17. 17. PubChem. Ciprofloxacin—PubChem [Internet] [cited 12 February 2020]. Available from: https://pubchem.ncbi.nlm.nih.gov/substance/375668987
  18. 18. Ezelarab H, Abbas SH, Hassan HA, Abuo-Rahma G. Recent updates of fluoroquinolones as antibacterial agents. Archiv der Pharmazie. 2018;351(9):e1800141. DOI: 10.1002/ardp.201800141
  19. 19. Grepafloxacin–PubChem [Internet] [cited 12 February 2020]. Available from: https://pubchem.ncbi.nlm.nih.gov/substance/318161242
  20. 20. Moxifloxacin—PubChem [Internet] [cited 12 February 2020]. Available from: https://pubchem.ncbi.nlm.nih.gov/substance/318161711
  21. 21. Quinolones and the Clinical Laboratory | HAI | CDC [Internet]. [cited 30 January 2020]. Available from: https://www.cdc.gov/hai/settings/lab/quinolones-clinical-laboratory.html
  22. 22. Correia S, Poeta P, Hébraud M, Capelo JL, Igrejas G. Mechanisms of quinolone action and resistance: Where do we stand? Journal of Medical Microbiology. 2017;66:551-559
  23. 23. Blanca-López N, Andreu I, Torrés Jaén MJ. Hypersensitivity reactions to quinolones. Current Opinion in Allergy and Clinical Immunology. 2011;11:285-291
  24. 24. Doña I, Moreno E, Pérez-Sánchez N, Andreu I, Hernández Fernandez de Rojas D, Torres MJ. Update on quinolone allergy. Current Allergy and Asthma Reports. 2017;17:1
  25. 25. Ashley RE, Dittmore A, McPherson SA, Turnbough CL, Neuman KC, Osheroff N. Activities of gyrase and topoisomerase IV on positively supercoiled DNA. Nucleic Acids Research. 2017;45(16):9611-9624
  26. 26. Nitiss JL. Roles of DNA topoisomerases in chromosomal replication and segregation. Advances in Pharmacology. 1994;29A:103-134
  27. 27. Collin F, Karkare S, Maxwell A. Exploiting bacterial DNA gyrase as a drug target: Current state and perspectives. Applied Microbiology and Biotechnology. 2011;92:479-497
  28. 28. Drlica K, Zhao X. DNA gyrase, topoisomerase IV, and the 4-quinolones. Microbiology and Molecular Biology Reviews. 1997;61(3):377-392
  29. 29. Ince D, Zhang X, Silver LC, Hooper DC. Dual targeting of DNA gyrase and topoisomerase IV: Target interactions of garenoxacin (BMS-284756, T-3811ME), a new desfluoroquinolone. Antimicrobial Agents and Chemotherapy. 2002;46(11):3370-3380
  30. 30. Reece RJ, Maxwell A, Wang JC. DNA gyrase: Structure and function. Critical Reviews in Biochemistry and Molecular Biology. 1991;26(3-4):335-375
  31. 31. Morais Cabral JH, Jackson AP, Smith CV, Shikotra N, Maxwell A, Liddington RC. Crystal structure of the breakage-Reunion domain of DNA gyrase. Nature. 1997;388(6645):903-906
  32. 32. Corbett KD, Schoeffler AJ, Thomsen ND, Berger JM. The structural basis for substrate specificity in DNA topoisomerase IV. Journal of Molecular Biology. 2005;351(3):545-561
  33. 33. Hooper DC, Jacoby GA. Mechanisms of drug resistance: Quinolone resistance. Annals of the New York Academy of Sciences. 2015;1354(1):12-31
  34. 34. Álvarez-Hernández DA, Garza-Mayén GS, Vázquez-López R. Quinolones. Nowadays perspectives and mechanisms of resistance. Revista chilena de infectología. 2015;32(5):499-504. DOI: 10.4067/S0716-10182015000600002
  35. 35. Hooper DC, Jacoby GA. Topoisomerase inhibitors: Fluoroquinolone mechanisms of action and resistance. Cold Spring Harbor Perspectives in Medicine. 2016;6(9):a025320. DOI: 10.1101/cshperspect.a025320
  36. 36. Wohlkonig A, Chan PF, Fosberry AP, Homes P, Huang J, Kranz M, et al. Structural basis of quinolone inhibition of type IIA topoisomerases and target-mediated resistance. Nature Structural & Molecular Biology. 2010;17(9):1152-1153
  37. 37. Laponogov I, Veselkov DA, Crevel IMT, Pan XS, Fisher LM, Sanderson MR. Structure of an “open” clamp type II topoisomerase-DNA complex provides a mechanism for DNA capture and transport. Nucleic Acids Research. 2013;41(21):9911-9923
  38. 38. Laponogov I, Sohi MK, Veselkov DA, Pan XS, Sawhney R, Thompson AW, et al. Structural insight into the quinolone-DNA cleavage complex of type IIA topoisomerases. Nature Structural & Molecular Biology. 2009;16(6):667-669
  39. 39. Yoshida H, Bogaki M, Nakamura M, Nakamura S. Quinolone resistance-determining region in the DNA gyrase gyrA gene of Escherichia coli. Antimicrobial Agents and Chemotherapy. 1990;34(6):1271-1272
  40. 40. Breines DM, Ouabdesselam S, Ng EY, Tankovic J, Shah S, Soussy CJ, et al. Quinolone resistance locus nfxD of Escherichia coli is a mutant allele of the parE gene encoding a subunit of topoisomerase IV. Antimicrobial Agents and Chemotherapy. 1997;41(1):175-179
  41. 41. Yoshida H, Bogaki M, Nakamura M, Yamanaka LM, Nakamura S. Quinolone resistance-determining region in the DNA gyrase gyrB gene of Escherichia coli. Antimicrobial Agents and Chemotherapy. 1991;35(8):1647-1650
  42. 42. Blanche F, Cameron B, Bernard FX, Maton L, Manse B, Ferrero L, et al. Differential behaviors of Staphylococcus aureus and Escherichia coli type II DNA topoisomerases. Antimicrobial Agents and Chemotherapy [Internet]. 1996;40(12):2714-2720 [cited 20 January 2020]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/9124828
  43. 43. Pan XS, Fisher LM. Targeting of DNA gyrase in Streptococcus pneumoniae by sparfloxacin: selective targeting of gyrase or topoisomerase IV by quinolones. Antimicrobial Agents and Chemotherapy [Internet]. 1997;41(2):471-474. Available from: http://www.ncbi.nlm.nih.gov/pubmed/9021211
  44. 44. Rahman T, Yarnall B, Doyle DA. Efflux drug transporters at the forefront of antimicrobial resistance. European Biophysics Journal. 2017;46:647-653
  45. 45. Ghai I, Ghai S. Understanding antibiotic resistance via outer membrane permeability. Infection and Drug Resistance. 2018;11:523-530
  46. 46. Grkovic S, Brown MH, Skurray RA. Regulation of bacterial drug export systems. Microbiology and Molecular Biology Reviews. 2002;66(4):671-701
  47. 47. Lekshmi M, Ammini P, Adjei JM, Sanford L, Shrestha U, Kumar S, et al. Modulation of antimicrobial efflux pumps of the major facilitator superfamily in Staphylococcus aureus. AIMS Microbiology. 2018;4(1):1-18
  48. 48. Sharma A, Gupta VK, Pathania R. Efflux pump inhibitors for bacterial pathogens: From bench to bedside. Indian Journal of Medical Research. 2019;149:129-145
  49. 49. JMA B, Richmond GE, Piddock LJV. Multidrug efflux pumps in Gram-negative bacteria and their role in antibiotic resistance. Future Microbiology [Internet]. 2014;9(10):1165-1177 [cited 24 January 2020]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25405886
  50. 50. Costa SS, Viveiros M, Amaral L, Couto I. Multidrug efflux pumps in Staphylococcus aureus: An update. The Open Microbiology Journal. 2013;7(1):59-71
  51. 51. Elbaz Y, Salomon T, Schuldiner S. Identification of a glycine motif required for packing in EmrE, a multidrug transporter from Escherichia coli. The Journal of Biological Chemistry. 2008;283(18):12276-12283
  52. 52. Bay DC, Turner RJ. Membrane composition influences the topology bias of bacterial integral membrane proteins. Biochimica et Biophysica Acta—Biomembranes. 2013;1828(2):260-270
  53. 53. Bay DC, Rommens KL, Turner RJ. Small multidrug resistance proteins: A multidrug transporter family that continues to grow [Internet]. Biochimica et Biophysica Acta—Biomembranes. 2008;1778:1814-1838 [cited 15 April 2020]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/17942072
  54. 54. Truong-Bolduc QC, Strahilevitz J, Hooper DC. NorC, a new efflux pump regulated by MgrA of Staphylococcus aureus. Antimicrobial Agents and Chemotherapy. 2006;50(3):1104-1107
  55. 55. Ubukata K, Itoh-Yamashita N, Konno M. Cloning and expression of the norA gene for fluoroquinolone resistance in Staphylococcus aureus. Antimicrobial Agents and Chemotherapy. 1989;33(9):1535-1539
  56. 56. Yoshida H, Bogaki M, Nakamura S, Ubukata K, Konno M. Nucleotide sequence and characterization of the Staphylococcus aureus norA gene, which confers resistance to quinolones. Journal of Bacteriology. 1990;172(12):6942-6949
  57. 57. Neyfakh AA, Borsch CM, Kaatz GW. Fluoroquinolone resistance protein NorA of Staphylococcus aureus is a multidrug efflux transporter. Antimicrobial Agents and Chemotherapy [Internet]. 1993;37(1):128-129 [cited 24 January 2020]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8431010
  58. 58. Costa SS, Sobkowiak B, Parreira R, Edgeworth JD, Viveiros M, Clark TG, et al. Genetic diversity of norA, coding for a main efflux pump of Staphylococcus aureus. Frontiers in Genetics. 2019;9:710. DOI: 10.3389/fgene.2018.00710
  59. 59. Ding Y, Onodera Y, Lee JC, Hooper DC. NorB, an efflux pump in Staphylococcus aureus strain MW2, contributes to bacterial fitness in abscesses. Journal of Bacteriology. 2008;190(21):7123-7129
  60. 60. Kaatz GW, Thyagarajan RV, Seo SM. Effect of promoter region mutations and mgrA overexpression on transcription of norA, which encodes a Staphylococcus aureus multidrug efflux transporter. Antimicrobial Agents and Chemotherapy. 2005;49(1):161-169
  61. 61. Truong-Bolduc QC, Hooper DC. The transcriptional regulators NorG and MgrA modulate resistance to both quinolones and β-lactams in Staphylococcus aureus. Journal of Bacteriology. 2007;189(8):2996-3005
  62. 62. Truong-Bolduc QC, Zhang X, Hooper DC. Characterization of NorR protein, a multifunctional regulator of norA expression in Staphylococcus aureus. Journal of Bacteriology [Internet]. 2003;(10):185, 3127-3138 [cited 25 January 2020] Available from: http://www.ncbi.nlm.nih.gov/pubmed/12730173
  63. 63. Truong-Bolduc QC, Bolduc GR, Okumura R, Celino B, Bevis J, Liao CH, et al. Implication of the NorB efflux pump in the adaptation of Staphylococcus aureus to growth at acid ph and in resistance to moxifloxacin. Antimicrobial Agents and Chemotherapy. 2011;55(7):3214-3219
  64. 64. Truong-Bolduc QC, Hsing LC, Villet R, Bolduc GR, Estabrooks Z, Florent Taguezem G, et al. Reduced aeration affects the expression of the NorB efflux pump of Staphylococcus aureus by posttranslational modification of MgrA. Journal of Bacteriology. 2012;194(7):1823-1834
  65. 65. Chen PR, Bae T, Williams WA, Duguid EM, Rice PA, Schneewind O, et al. An oxidation-sensing mechanism is used by the global regulator MgrA in Staphylococcus aureus. Nature Chemical Biology. 2006;2(11):591-595
  66. 66. Truong-Bolduc QC, Dunman PM, Eidem T, Hooper DC. Transcriptional profiling analysis of the global regulator NorG, a GntR-like protein of Staphylococcus aureus. Journal of Bacteriology. 2011;193(22):6207-6214
  67. 67. Huang J, O’Toole PW, Shen W, Amrine-Madsen H, Jiang X, Lobo N, et al. Novel chromosomally encoded multidrug efflux transporter MdeA in Staphylococcus aureus. Antimicrobial Agents and Chemotherapy. 2004;48(3):909-917
  68. 68. Yamada Y, Shiota S, Mizushima T, Kuroda T, Tsuchiya T. Functional gene cloning and characterization of MdeA, a multidrug efflux pump from Staphylococcus aureus. Biological & Pharmaceutical Bulletin. 2006;29(4):801-804
  69. 69. Yamada Y, Hideka K, Shiota S, Kuroda T, Tsuchiya T. Gene cloning and characterization of SdrM, a chromosomally-encoded multidrug efflux pump, from Staphylococcus aureus. Biological and Pharmaceutical Bulletin [Internet]. 2006;29(3):554-556 [cited 26 January 2020] Available from: http://joi.jlc.jst.go.jp/JST.JSTAGE/bpb/29.554?from=CrossRef
  70. 70. Nakaminami H, Noguchi N, Sasatsu M. Fluoroquinolone efflux by the plasmid-mediated multidrug efflux pump QacB variant QacBIII in Staphylococcus aureus. Antimicrobial Agents and Chemotherapy. 2010;54(10):4107-4111
  71. 71. Dabul ANG, Avaca-Crusca JS, Van Tyne D, Gilmore MS, Camargo ILBC. Resistance in in vitro selected tigecycline-resistant methicillin-resistant Staphylococcus aureus sequence type 5 is driven by mutations in mepR and mepA genes. Microbial Drug Resistance. 2018;24(5):519-526
  72. 72. McAleese F, Petersen P, Ruzin A, Dunman PM, Murphy E, Projan SJ, et al. A novel MATE family efflux pump contributes to the reduced susceptibility of laboratory-derived Staphylococcus aureus mutants to tigecycline. Antimicrobial Agents and Chemotherapy. 2005;49(5):1865-1871
  73. 73. Kaatz GW, McAleese F, Seo SM. Multidrug resistance in Staphylococcus aureus due to overexpression of a novel multidrug and toxin extrusion (MATE) transport protein. Antimicrobial Agents and Chemotherapy. 2005;49(5):1857-1864
  74. 74. Nikaido H. Structure and mechanism of RND-type multidrug efflux pumps. Advances in Enzymology and Related Areas of Molecular Biology. 2010;77(1):1-60
  75. 75. Anes J, McCusker MP, Fanning S, Martins M. The ins and outs of RND efflux pumps in Escherichia coli. Frontiers in Microbiology. 2015;6:587. DOI: 10.3389/fmicb.2015.00587
  76. 76. Venter H, Mowla R, Ohene-Agyei T, Ma S. RND-type drug efflux pumps from Gram-negative bacteria: Molecular mechanism and inhibition. Frontiers in Microbiology. 2015;6:377. DOI: 10.3389/fmicb.2015.00377
  77. 77. Kim EH, Nies DH, McEvoy MM, Rensing C. Switch or funnel: How RND-type transport systems control periplasmic metal homeostasis. Journal of Bacteriology. 2011;193:2381-2387
  78. 78. Tikhonova EB, Zgurskaya HI. AcrA, AcrB, and TolC of Escherichia coli form a stable intermembrane multidrug efflux complex. The Journal of Biological Chemistry. 2004;279(31):32116-32124
  79. 79. Pos KM. Drug transport mechanism of the AcrB efflux pump. Biochimica et Biophysica Acta, Proteins and Proteomics. 2009;1794:782-793
  80. 80. Rosenberg EY, Ma D, Nikaido H. AcrD of Escherichia coli is an aminoglycoside efflux pump. Journal of Bacteriology. 2000;182(6):1754-1756
  81. 81. Elkins CA, Nikaido H. Substrate specificity of the RND-type multidrug efflux pumps AcrB and AcrD of Escherichia coli is determined predominantly by two large periplasmic loops. Journal of Bacteriology [Internet]. 2002;184(23):6490-6498 [cited 31 January 2020]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12426336
  82. 82. Nishino K, Yamaguchi A. Role of histone-like protein H-NS in multidrug resistance of Escherichia coli. Journal of Bacteriology. 2004;186(5):1423-1429
  83. 83. Baranova N, Nikaido H. The baeSR two-component regulatory system activates transcription of the yegMNOB (mdtABCD) transporter gene cluster in Escherichia coli and increases its resistance to novobiocin and deoxycholate. Journal of Bacteriology [Internet]. 2002;184(15):4168-4176 [cited 31 January 2020]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12107134
  84. 84. Nagakubo S, Nishino K, Hirata T, Yamaguchi A. The putative response regulator BaeR stimulates multidrug resistance of Escherichia coli via a novel multidrug exporter system, MdtABC. Journal of Bacteriology [Internet]. 2002;184(15):4161-4167 [cited 27 January 2020]. Available from: http://www.ncbi.nlm.nih.gov/pubmed/12107133
  85. 85. Hirakawa H, Inazumi Y, Senda Y, Kobayashi A, Hirata T, Nishino K, et al. N-acetyl-D-glucosamine induces the expression of multidrug exporter genes, mdtEF, via catabolite activation in Escherichia coli. Journal of Bacteriology. 2006;188(16):5851-5858
  86. 86. Zhang Y, Xiao M, Horiyama T, Zhang Y, Li X, Nishino K, et al. The multidrug efflux pump MdtEF protects against nitrosative damage during the anaerobic respiration in Escherichia coli. The Journal of Biological Chemistry. 2011;286(30):26576-26584
  87. 87. Du D, Wang Z, James NR, Voss JE, Klimont E, Ohene-Agyei T, et al. Structure of the AcrAB-TolC multidrug efflux pump. Nature. 2014;509(7501):512-515
  88. 88. Yu EW, McDermott G, Zgurskaya HI, Nikaido H, Koshland DE. Structural basis of multiple drug-binding capacity of the AcrB multidrug efflux pump. Science. 2003;300(5621):976-980
  89. 89. Murakami S, Nakashima R, Yamashita E, Yamaguchi A. Crystal structure of bacterial multidrug efflux transporter AcrB. Nature. 2002;419(6907):587-593
  90. 90. Poole K. Efflux-mediated multiresistance in Gram-negative bacteria. Clinical Microbiology and Infection. 2004;10:12-26
  91. 91. Lau SY, Zgurskaya HI. Cell division defects in Escherichia coli deficient in the multidrug efflux transporter AcrEF-TolC. Journal of Bacteriology. 2005;187(22):7815-7825
  92. 92. Poole K. Efflux-mediated resistance to fluoroquinolones in gram-negative bacteria. Antimicrobial Agents and Chemotherapy. 2000;44:2233-2241
  93. 93. Nishino K, Nikaido E, Yamaguchi A. Regulation of multidrug efflux systems involved in multidrug and metal resistance of Salmonella enterica serovar typhimurium. Journal of Bacteriology. 2007;189(24):9066-9075
  94. 94. Nagakubo S, Nishino K, Hirata T, Yamaguchi A. The putative response regulator BaeR stimulates multidrug resistance of Escherichia coli via a novel multidrug exporter system, MdtABC. Journal of Bacteriology [Internet]. 2002;184(15):4161-4167 [cited 31 January 2020] Available from: http://www.ncbi.nlm.nih.gov/pubmed/12107133
  95. 95. Choudhury D, Ghosh A, Chanda DD, Das TA, Choudhury MD, Paul D, et al. Premature termination of MexR leads to overexpression of MexAB-OprM efflux pump in Pseudomonas aeruginosa in a tertiary referral hospital in India. PLoS One. 2016;11(2):1
  96. 96. Li XZ, Zhang L, Poole K. Interplay between the MexA-MexB-OprM multidrug efflux system and the outer membrane barrier in the multiple antibiotic resistance of Pseudomonas aeruginosa. The Journal of Antimicrobial Chemotherapy. 2000;45(4):433-436
  97. 97. Masuda N, Gotoh N, Ohya S, Nishino T. Quantitative correlation between susceptibility and OprJ production in NfxB mutants of Pseudomonas aeruginosa. Antimicrobial Agents and Chemotherapy. 1996;40(4):909-913
  98. 98. Richardot C, Juarez P, Jeannot K, Patry I, Plésiat P, Llanes C. Amino acid substitutions account for most mexS alterations in clinical nfxC mutants of Pseudomonas aeruginosa. Antimicrobial Agents and Chemotherapy. 2016;60(4):2302-2310
  99. 99. Mine T, Morita Y, Kataoka A, Mizushima T, Tsuchiya T. Expression in Escherichia coli of a new multidrug efflux pump, MexXY, from Pseudomonas aeruginosa. Antimicrobial Agents and Chemotherapy. 1999;43(2):415-417
  100. 100. Hocquet D, Vogne C, El Garch F, Vejux A, Gotoh N, Lee A, et al. MexXy-OprM efflux pump is necessary for adaptive resistance of Pseudomonas aeruginosa to aminoglycosides. Antimicrobial Agents and Chemotherapy. 2003;47(4):1371-1375
  101. 101. Lin J, Akiba M, Sahin O, Zhang Q . CmeR functions as a transcriptional repressor for the multidrug efflux pump CmeABC in Campylobacter jejuni. Antimicrobial Agents and Chemotherapy. 2005;49(3):1067-1075
  102. 102. Yao H, Shen Z, Wang Y, Deng F, Liu D, Naren G, et al. Emergence of a potent multidrug efflux pump variant that enhances Campylobacter resistance to multiple antibiotics. MBio. 2016;7(5):1
  103. 103. Damier-Piolle L, Magnet S, Brémont S, Lambert T, Courvalin P. AdeIJK, a resistance-nodulation-cell division pump effluxing multiple antibiotics in Acinetobacter baumannii. Antimicrobial Agents and Chemotherapy. 2008;52(2):557-562
  104. 104. Ruzin A, Keeney D, Bradford PA. AdeABC multidrug efflux pump is associated with decreased susceptibility to tigecycline in Acinetobacter calcoaceticus-Acinetobacter baumannii complex. The Journal of Antimicrobial Chemotherapy. 2007;59(5):1001-1004
  105. 105. Coyne S, Courvalin P, Périchon B. Efflux-mediated antibiotic resistance in Acinetobacter spp. Antimicrobial Agents and Chemotherapy. 2011;55:947-953
  106. 106. Coyne S, Rosenfeld N, Lambert T, Courvalin P, Perichon B. Overexpression of resistance-nodulation-cell division pump AdeFGH confers multidrug resistance in Acinetobacter baumannii. Antimicrobial Agents and Chemotherapy [Internet]. 2010;54(10):4389-4393 [cited 08 March 2019] Available from: http://www.ncbi.nlm.nih.gov/pubmed/20696879
  107. 107. Zhang L, Li XZ, Poole K. SmeDEF multidrug efflux pump contributes to intrinsic multidrug resistance in Stenotrophomonas maltophilia. Antimicrobial Agents and Chemotherapy. 2001;45(12):3497-3503
  108. 108. Lomovskaya O, Lewis K, Matin A. EmrR is a negative regulator of the Escherichia coli multidrug resistance pump emrAB. Journal of Bacteriology. 1995;177(9):2328-2334
  109. 109. Bohn C, Bouloc P. The Escherichia coli cmlA gene encodes the multidrug efflux pump Cmr/MdfA and is responsible for isopropyl-β-D-thiogalactopyranoside exclusion and spectinomycin sensitivity. Journal of Bacteriology. 1998;180(22):6072-6075
  110. 110. Zhao Y, Heng J, Zhao Y, Liu M, Liu Y, Fan J, et al. Substrate-bound structure of the E. coli multidrug resistance transporter MdfA. Cell Research. 2015;25(9):1060-1073
  111. 111. Nestorovich EM, Sugawara E, Nikaido H, Bezrukov SM. Pseudomonas aeruginosa porin OprF. Properties of the channel. The Journal of Biological Chemistry. 2006;281(24):16230-16237
  112. 112. Nikaido H, Nikaido K, Harayama S. Identification and characterization of porins in Pseudomonas aeruginosa. The Journal of Biological Chemistry. 1991;266(2):770-779
  113. 113. Martínez-Martínez L, Pascual A, Jacoby GA. Quinolone resistance from a transferable plasmid. Lancet. 1998;351(9105):797-799
  114. 114. Jacoby GA. Mechanisms of resistance to quinolones. Clinical Infectious Diseases [Internet]. 2005;41(Supplement_2):S120-S126. [cited 17 January 2020]. Available from: http://academic.oup.com/cid/article/41/Supplement_2/S120/307501/Mechanisms-of-Resistance-to-Quinolones
  115. 115. Tran JH, Jacoby GA. Mechanism of plasmid-mediated quinolone resistance. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(8):5638-5642
  116. 116. Jacoby GA, Walsh KE, Mills DM, Walker VJ, Oh H, Robicsek A, et al. qnrB, another plasmid-mediated gene for quinolone resistance. Antimicrobial Agents and Chemotherapy. 2006;50(4):1178-1182
  117. 117. Hata M, Suzuki M, Matsumoto M, Takahashi M, Sato K, Ibe S, et al. Cloning of a novel gene for quinolone resistance from a transferable plasmid in Shigella flexneri 2b. Antimicrobial Agents and Chemotherapy. 2005;49(2):801-803
  118. 118. Robicsek A, Strahilevitz J, Jacoby GA, Macielag M, Abbanat D, Chi HP, et al. Fluoroquinolone-modifying enzyme: A new adaptation of a common aminoglycoside acetyltransferase. Nature Medicine. 2006;12(1):83-88
  119. 119. Yamane K, Wachino JI, Suzuki S, Kimura K, Shibata N, Kato H, et al. New plasmid-mediated fluoroquinolone efflux pump, QepA, found in an Escherichia coli clinical isolate. Antimicrobial Agents and Chemotherapy. 2007;51(9):3354-3360
  120. 120. Périchon B, Courvalin P, Galimand M. Transferable resistance to aminoglycosides by methylation of G1405 in 16S rRNA and to hydrophilic fluoroquinolones by QepA-mediated efflux in Escherichia coli. Antimicrobial Agents and Chemotherapy. 2007;51(7):2464-2469
  121. 121. Hansen LH, Jensen LB, Sørensen HI, Sørensen SJ. Substrate specificity of the OqxAB multidrug resistance pump in Escherichia coli and selected enteric bacteria. Journal of Antimicrobial Chemotherapy [Internet]. 2007;60(1):145-147. [cited 30 January 2020]. Available from: http://academic.oup.com/jac/article/60/1/145/731003/Substrate-specificity-of-the-OqxAB-multidrug

Written By

Sandra Georgina Solano-Gálvez, María Fernanda Valencia-Segrove, María José Ostos Prado, Ana Berenice López Boucieguez, Diego Abelardo Álvarez-Hernández and Rosalino Vázquez-López

Submitted: 27 September 2019 Reviewed: 16 April 2020 Published: 03 July 2020